Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Attosecond chirp effect on the transient absorption spectrum of laser-dressed helium atom

Open Access Open Access

Abstract

We theoretically investigate the attosecond transient absorption spectrum of helium atom in the presence of an infrared-dressed laser pulse upon scanning their relative delay, with the particular emphasis on the chirp effect of the attosecond pulse. By numerically solving the fully three-dimensional time-dependent Schrödinger equation, we identify the attoscecond chirp can induce the temporal shift of the absorption spectrogram along the delay axis. Additionally, it is found that the extent of the temporal shift is dependent on both the position of the absorption line and the infrared pulse wavelength, which is well confirmed and reproduced by a three-level model. Moreover, we demonstrate that the observed features can be quantitatively explained in terms of the indirect two-photon absorption processes through some virtual states. This effect might provide a way to measure the chirp of attosecond pulse in an all-optical way.

© 2017 Optical Society of America

1. Introduction

Since the first attosecond pulse was obtained experimentally, various attosecond spectroscopies have been reported and implemented to probe ultrafast electron dynamics in matters [1–4]. As one of the outstanding representatives, attosecond transient absorption spectroscopy (ATAS) provides an unprecedented temporal resolution for real-time observation of electron dynamics on its natural time scale [5, 6]. ATAS usually arises from the interaction between an attosecond extreme-ultraviolet (XUV) pulse and a matter target dressed by a few-cycle infrared (IR) pulse, and records the spectral change of the XUV pulse passing through the target versus the delay between these two pulses, which reveals the transient dynamics in the interaction.

The first successful application of the ATAS is the measurement of the valence electron motion in krypton ions generated by the controlled few-cycle IR pulse [6], which motivates a great deal of experimental investigations based on the ATAS technique to achieve the ultrafast measurement from the gas phase to the condensed matter. It has been demonstrated that the ATAS is successfully used in numerous areas, such as revealing the interference of transiently bound electron wave packets [7], probing the optical Stark shift of the laser-dressed 1s3p state of helium atom [8], reconstructing the electronic and nuclear dynamics in hydrogen molecules [9], resolving the ultrafast band-gap dynamics in silicon in real time [10], and others [11–16]. Besides these advances in experiment, theoretical calculation of the ATAS was realized by a time-frequency method [17]. Using this method, the nuclear-motion effects of molecules in the ATAS were analyzed [18] and the quantum beats were first observed in the simulated absorption spectrum of helium [19]. Theoretical development enables the further investigation of the rich dynamics of ATAS [20–25].

To date, theoretical researches on ATAS are all focused on the use of a chirp-free XUV pulse to pump (probe) an excited electron wave packet which is probed (pumped) by a few-cycle IR laser [18, 21, 24], and the chirp characteristic of attosecond pulse has been ignored. Nevertheless, chirp is difficult to eliminate in attosecond pulse generated by high harmonics due to the intrinsically laser-intensity-dependent dipole [26]. Moreover, recent studies have demonstrated that chirped laser pulses possess selectivity of controlling the population transfer in the process of two-photon double ionization of atoms [27, 28]. Taken this into consideration, we guess that chirped attosecond pulses may have influence on ATAS which contains a similar two-photon process as pump-probe absorption measurements. What the additional effect remains to be addressed.

In this work, we report the first study of attosecond pulse chirp effects on transient absorption spectrum. Firstly, we introduce the chirped Gaussian attosecond pulses with different linear chirp rates instead of chirp-free pulse to calculate the ATAS of laser-dressed helium atom by numerically solving the fully three-dimensional ti spectra exhibit almost the same structure for different chirps, obvious relative translation between them along the delay axis is observed, which gives evidence of the existence of chirp effect on ATAS. Secondly, considering that the 3D-TDSE calculation is time consuming and difficult to reveal inherent physical mechanism, we further employ the three-level model which is a more efficient method to analyze the contribution of specific electronic states [19, 29]. According to this model, we have successfully reproduced the chirp effect and through the analysis of the IR-field-dressed eigenstates, we find that chirp effect is associated with indirect two-photon absorption pathways involving the absorption of an XUV photon and then the absorption or emission of an IR photon via a set of the virtual states, so that the electron can transfer from the ground state to these eigenstates.

2. Theoretical model and numerical method

The ATAS of helium atom in the presence of the dressed IR pulse is calculated by numerically solving the fully 3D-TDSE under the single active electron approximation. This treatment has been widely used in many previous works and proved effective in reproducing main features of ATAS in helium [14, 17, 23]. In the Cartesian spherical coordinate, the fully 3D-TDSE is expressed as

itΨ(r,t)=[121r2r+l^22r2+V(r)+VI(r,t)]Ψ(r,t),
where l^2 is the square of the orbit angular momentum operator, V(r) is an effective Coulomb potential with spherical symmetry, given by [30]
V(r)=1.0+1.231e0.662r1.325re1.236r0.231e0.48rr,
and VI(r,t) describes the interaction Hamiltonian for the helium atom irradiated by the external pulse field.

The ATAS is always implemented upon scanning the relative delay of the IR pulse and the XUV pulse. We assume that the two pulses both have the polarization direction along the z axis. In the length gauge and dipole approximation, the VI(r,t) can be described as VI(r,t)=zE(t), where the total electric field is related with the corresponding vector potential A(t) as E(t)=ddtA(t). Here the total vector potential of two pulses can be written as A(t)=AX(ttD)+AIR(t) with the delay tD. The time-dependent vector potentials of a chirped XUV Gaussian pulse and a chirp-free IR Gaussian pulse can be written respectively as [31]

AX(ttD)=Re(i1ωXIX1iξexp{i[ωX(ttD)+φX]2ln2(ttD)2τX2(1iξ)}),
and
AIR(t)=Re(i1ωIRIIRexp{i[ωIRt+φIR]2ln2t2τIR2}),
where ωX(ωIR) is the central carrier frequency, φX(φIR) is the carrier envelope phase (CEP), IX(IIR) and τX(τIR) are the peak intensity and duration (full width at half maximum) of the transform-limited pulse respectively, and ξ is the dimensionless linear chirp rate of XUV attosecond pulse. Equation (3) describes the realistic situation that while the duration and peak intensity of the pulse depend on the chirp rate ξ, the spectral profile and energy of the pulse do not depend on ξ.

A detailed description of the numerical method used to solve Eq. (1) is reported in [32]. In brief, the time-dependent wave function Ψ(r,t) is firstly expanded into a series of spherical harmonics. As a result, Eq. (1) can reduce to a set of coupled equations between the different angular quantum numbers for the radial wave function, which is further discretized by the finite-element discrete variable representation (FE-DVR) method with the advantage of providing block-diagonal sparse matrix representation of kinetic operator and the diagonal matrix representation of the effective coulomb potential. The temporal evolution of the wave function is carried out by the Arnoldi-Lanczos algorithm.

After the fully 3D-TDSE is solved, the transient absorption spectrum can be achieved by calculating the single-atom frequency-dependent response function [17]

S˜(ω,tD)=2Im[d˜(ω,tD)ε˜*(ω,tD)],ω>0,
where tD is the pump-probe delay, d˜(ω,tD) and ε˜*(ω,tD) are respectively the Fourier transforms of the time-dependent dipole moment d(t) and the total driving field E(t) for a given pump-probe delay. We calculate the dipole spectrum d˜(ω) via the time-dependent dipole acceleration a(t) which is obtained by the Ehrenfest theorem. The dipole spectrum is then given as d˜(ω)=a˜(ω)/ω2. Before calculating a˜(ω) by Fourier transforms, a(t) is multiplied by a finite time window of 65 fs in order to eliminate the high-frequency numerical noise in the process of Fourier transform. S˜(ω,tD) represents the excitation probability per unit frequency at a given frequency. For position frequencies, ωS˜(ω) equals the energy gained or lost per unit frequency by the atom, meaning that positive or negative values of S˜(ω) express absorption or emission of light at frequency ω.

3. Results and discussion

In order to reveal the chirp effect of attosecond pulse on the transient absorption of IR-dressed helium atom, we have firstly calculated the single-atom response function S˜(ω,tD) using the fully 3D-TDSE model upon scanning the relative delay between the IR pulse and the XUV attosecond pulse.

Figure 1 shows the calculated ATASs in helium for three different chirp rates of the XUV pulses: (a) ξ=4, (b) ξ=0 and (c) ξ=4. Note that the negative (positive) delay represents the XUV pulses arrive before (after) the IR pulse. In the calculation, the XUV pulses have a central photon energy of 24 eV, a transform-limited duration of 300 as and a peak intensity of 1×1012W/cm2. These pulses have a bandwidth of 6 eV, which can promote the electron from the ground state to all singly excited states below the single ionization threshold (Ip) and to low-energy continuum states of helium atom [23]. The IR pulse with the 800 nm wavelength has the duration of 10 fs and a relatively moderate peak intensity of 1×1013W/cm2. We point out that a relatively moderate IR intensity of 1×1013W/cm2 is a proper to make the chirp of attosecond pulse readily observable in absorption spectra. The weaker (~1×1012W/cm2) or the stronger (~1×1014W/cm2) IR intensity will induce too low contrast of absorption signals or the severe shift and splitting of energy levels, which prevents the possible observation in the experiment.

 figure: Fig. 1

Fig. 1 Calculated single-atom response function S˜(ω,tD) using the fully 3D-TDSE model upon scanning the relative delay between the IR pulse and the attosecond pulse with three different chirp rates: (a) ξ=4, (b) ξ=0 and (c) ξ=4.

Download Full Size | PDF

One can see several similar features regardless of the attosecond chirp effect in Figs. 1(a)-1(c) as follow: (i) For large positive delays, there are the narrow absorption lines induced only by the XUV pulse, representing the excitation of the electron from the 1s ground state to all the np excited states by single-photon transition. (ii) For large negative delays, a modulation of absorption following hyperbolic sidebands is referred to as perturbed free-induction decay [33]. (iii) For the temporal overlap region of the two pulses, the modulation of absorption with the period of half-IR-cycle and the light-induced structure (LIS) are present between the 2p and 3p resonant lines, which is firstly observed experimentally in 2012 [12].

Although Figs. 1(a)-1(c) exhibit almost the same structure, we can clearly reveal the attosecond chirp effect by extracting the absorption lines from Fig. 1 at three representative states: 2p (21.05 eV), 4p (23.73 eV) and a low-energy continuum state (25.55 eV). Note that the ground state energy is already set to zero. Figures 2(a)-2(c) respectively show the delay-dependent absorption signals near 21.05 eV, 23.73 eV and 25.55 eV, calculated as the integral of the response function around each state with a spectral width of 0.5 eV, for three different attosecond chirp rates. The primary feature included in the Figs. 2(a)-2(c) is the temporal shift of absorption curves with varying the chirp rate of the attosecond pulse, together with the shift distance dependent on the absorption line energy of the observation.

 figure: Fig. 2

Fig. 2 Delay-dependent absorption signals near (a) 2p (21.05 eV), (b) 4p (23.73 eV) and (c) the low-energy continuum state (25.55 eV), respectively for three different attosecond chirp rates: ξ=4 (red-solid line), ξ=0 (blue-dashed line) and ξ=4 (black-dot line).

Download Full Size | PDF

The fully 3D-TDSE calculation is time consuming and difficult to identify the underling mechanism due to the complete consideration of all electronic states and interaction dynamics. We further employ the N-level model which is an efficient and compact method [19, 29], to analyze the contribution of specific electronic states to the attosecond chirp effect on the ATAS.

As an example, we mainly focus on the absorption signal near the 2p state by choosing three energy levels composed of the 1s, 2p and 3s states. In general, both the IR and XUV pulses can lead to the transition among these three states. In order to separate the contribution of the XUV and IR pulses, we artificially switch off the transition between 1s and 2p induced by the IR pulse, and the transition between 2p and 3s induced by the XUV pulse, which means that the XUV pulse only excites 1s to 2p and the IR pulse only excites 2p to 3s. In this situation, we calculate the attosecond chirp effect by analyzing the absorption signals near the 2p state as a function of the delay under three different attosecond chirp rates: ξ=4 (red-solid), ξ=0 (blue-dashed) and ξ=4 (black-dot), as shown in Fig. 3. One can see that the 2p absorption signals in Fig. 3 maintain significant features compared with the fully 3D-TDSE result shown in Fig. 2(a), including the same modulation period of 1.33 fs and nearly the same phase shift of peak positons with respect to the chirp rate ξ. This comparison demonstrates the three-level model instead of the full 3D-TDSE calculation is sufficient and valid to capture the important characteristics of the chirp-dependent ATAS.

 figure: Fig. 3

Fig. 3 Absorption signals near the 2p state (21.05 eV) in the case of three different chirp rates: ξ=4 (red-solid line), ξ=0 (blue-dashed line) and ξ=4 (black-dot line), calculated by three-level model. Other simulation parameters are the same as Fig. 1.

Download Full Size | PDF

For quantitatively identifying the chirp effect, we can define the delay offset Δ(ξ) for a given chirp rate ξ relative to the chirp-free case as Δ(ξ)=tD(ξ)tD(0), where tD(ξ) corresponds to certain absorption peaks selected by the arrows in Fig. 3. We use the value of delay offset Δ(4) to estimate the influences of the IR duration and the additional state couplings. We find that for both 5-fs and 10-fs IR pulses, the additional couplings beyond the 3-level model will result in the variation of Δ(4) less than 0.04 fs. It is expected that the contribution from the other Rydberg states such as 2s and 3d coupling to 2p state will become weaker with increasing the IR duration. As a result, the chirp-dependent absorption spectrum is mainly determined by the 3-level model for long IR pulse. If this effect is used to determine the chirp of attosecond pulse, the accuracy of the retrieved chirp can been estimated by δξ=Δ5(ξ0)Δ3(ξ0)dΔ3(ξ)dξ|ξ=ξ0, where Δ5(ξ0) (Δ3(ξ0)) is the delay offset for a given chirp rate of ξ0 calculated by 5-level model (3-level model). Here the 5-level model consists of 1s, 2p, 3s, 2s, and 3d states. Since dΔ3(ξ)/dξ is approximately equal to 0.066 fs, δξ is less than 0.04/0.066=0.6061.

Since the wavelength of the IR pulse is an important parameter in the strong-field process, we have further investigated how the IR wavelength affects the Δ(ξ) . Figure 4 shows Δ(ξ) as a function of the IR wavelength at five different chirp rates: ξ=4 (red-pluses), ξ=2 (green-stars), ξ=0 (blue-circles), ξ=2 (cyan-crosses) and ξ=4 (black-squares). It is found that Δ(ξ) is linearly dependent on the chirp rate ξ and also changes with the IR wavelength. For the positive (negative) attosecond chirp rate ξ, Δ(ξ) increases (decreases) with increasing IR wavelength until reaching the saturation value in the long wavelength limit. Moreover, Δ(ξ) can change its sign at 420 nm position for all chirp values.

 figure: Fig. 4

Fig. 4 The IR wavelength-dependent offset Δ(ξ)=tD(ξ)tD(0) in the case of five different chirp rates: ξ=4 (red-pluses), ξ=2 (green-stars), ξ=0 (blue-circles), ξ=2 (cyan-crosses) and ξ=4 (black-squares) presents the fully 3D-TDSE results. Other parameters are consistent with Fig. 3.

Download Full Size | PDF

For a deep insight into the chirp effect on ATAS, we propose an indirect two-photon process based on the three-level model to interpret the above simulation results. The intense IR pulse can lead to the energy shift and the state mixing of the field-free energy eigenstates. To simplify the analysis and highlight the physical picture, we only consider the field-free eigenstates 2p and 3s affected by the IR pulse, which will result in the generation of two IR-dressed instantaneous eigenstates, denoted as S and S+ . Their energy eigenvalues can be expressed respectively as

E=E2p+E3s(E2pE3s)2+4d232EIR2(t)2,
E+=E2p+E3s+(E2pE3s)2+4d232EIR2(t)2,
where E2p(E3s) represents the energy of the 2p(3s) state, EIR(t) is the instantaneous amplitude of the IR electric field. For all parameters considered in our work, we can estimate the energy shift of dressed eigenstates less than 0.14 eV.

Both S and S+ states are the mixture of the 2p and 3s states, in which the proportion of 3s state in S and S+ are given respectively by

η=2d232EIR2(t)(E2pE3s)2+4d232EIR2(t)(E2pE3s)(E2pE3s)2+4d232EIR2(t),
η+=2d232EIR2(t)(E2pE3s)2+4d232EIR2(t)+(E2pE3s)(E2pE3s)2+4d232EIR2(t).
Here, η(η+) is almost proportional (inversely proportional) to the IR intensity and reaches the highest (lowest) value 0.07 (0.93) when EIR2(t)=1×1013W/cm2.

The schematic of the indirect two-photon process is shown in Fig. 5, where the transition from the ground state 1s (E1s=0eV) to S eigenstate can occur through an alternative indirect ωωIR pathway associated with the virtual state S+ with the energy ΔE+ωIR, where ΔE=EE1sE2pE1s. Because the mixed S eigenstate include the 3s state component, this two-photon process is permitted without violating the dipole transition selection rule. The transition process for the electron excited from the ground state to the S state requires that the helium atom gains the net energy ΔE21.05eV. This can be achieved in the indirect way by simultaneously absorbing one XUV photon and radiating one IR photon. Consequently, the response function S˜(ΔE,tD) describing the absorption line near the 2p (E2p=21.05eV) state can become the maximum value only when the XUV photon with the energy ΔE+ωIR reaches the maximal intensity. The chirp in the attosecond pulse leads to the time-dependent instantaneous frequency, so that the moment tm at which the ΔE+ωIR photon component reaches the maximal intensity is different for different chirp rates.

 figure: Fig. 5

Fig. 5 Schematic of the three-level model including the 1s, 2p and 3s states. S and S+ are eigenstates for the 2p and 3s states dressed by the intense IR pulse. The indirect two-photon pathway from the ground state 1s to the final state S through the virtual state S+ is also indicated.

Download Full Size | PDF

We can acquire the chirp-dependent moment tm(ξ) by performing time-frequency analysis to the XUV pulses for different chirp rates, as presented in Fig. 6(a). For a fixed energy E and along the horizontal direction (t axis) of Fig. 6(a), we can obtain a peak value which is related with the certain time T. This means that the E photon component involved in the chirped attosecond pulse has the strongest intensity at the time T. We can plot E versus T for five different chirp rates, as shown in Fig. 6(b). The proposed two-photon absorption process implies that the efficient transition can occur when the XUV photon possesses the energy ωXUV=ΔE+ωIR, whose position is denoted by the black-dashed horizontal line in Fig. 6(b). Consequently, the t-coordinate of the cross points of the horizontal line and those titled lines correspond to tm(ξ,ωXUV) (ξ=4,2,0,2,4), as indicated in Fig. 6(b). It follows that the offset Δ(ξ,ωXUV) for the S absorption peak can be given by Δ(ξ,ωXUV)=tm(0,ωXUV)tm(ξ,ωXUV).

 figure: Fig. 6

Fig. 6 (a) Time-frequency spectrogram of the XUV pulses at different chirp rates: ξ=4,2,0,2,4 respectively. (b) Time-dependent photon energy which has the maximal intensity, extracted from panel (a) for five different chirp rates.

Download Full Size | PDF

According to ωIR(eV)1240/λ(nm) (λ is IR wavelength), we can further obtain the IR wavelength-dependent offset Δ for different chirp rates ξ, as shown in Fig. 7. The nice agreement between Fig. 7 and Fig. 4 supports the two-photon absorption picture. It is worth mentioning that there is a special IR wavelength 420nm for which the temporal offset Δ is equal to zero for all values of the chirp. We point out that the 420nm wavelength is related with the central photon energy of the XUV pulse. As already stated above, the required XUV photon energy is close to ωXUV=21.05+ωIR eV and denoted by the dashed horizontal line in Fig. 6 (b). In order to make the delay offsets Δ is equal to zero for all values of the chirp ξ, the ωXUV horizontal line through the dependence on the ωIR should meet the S point shown in Fig. 6 (b). It follows that the special IR wavelength which leads to Δ(ξ)=0 can be given by λ(nm)1240/(ωS21.05eV). Here ωS is the energy corresponding to S point, representing the intensity of the ωS photon component will reach the maximum at the same moment for all chirp rates ξ. The ωS is always equal to the central photon energy of the XUV pulse. In our simulation, the central photon energy of the XUV pulse is chosen as 24 eV. We can therefore obtain the special wavelength λ(nm)1240/(2421.05eV)=420nm.

 figure: Fig. 7

Fig. 7 The IR wavelength-dependent offset Δ(ξ) for five different chirp rates: ξ=4,2,0,2,4 calculated by an analytical model based on the two-photon absorption process. Other parameters are consistent with Fig. 4.

Download Full Size | PDF

The indirect two-photon absorption process can also be used to explain the chirp effect near the 4p state (E4p=23.73eV) shown in Fig. 2(b) and the low-energy continuum state (Ec=25.55eV) shown in Fig. 2(c).

For the absorption signal near the 4p state, the electron can be excited from the ground state to IR-dressed 4p state, whose energy is close to field-free 4p state, by simultaneously absorbing one XUV photon and one 800 nm IR photon (XUV + IR process). In this case, we can replace ωXUV in the above analysis with the required photon energy ωXUVE4pE1sωIR=22.18eV, whose position corresponds to the downward shift of the black-dashed horizontal line shown in Fig. 6(b), inducing the increase of the interval of the adjacent intersection. As a consequence, the offset Δ(ξ) becomes larger than the one in the 2p state case.

For the absorption signal near the continuum state Ec=25.55eV, the similar XUV + IR process requires the XUV photon energy ωXUVEcE1sωIR=24eV, whose position corresponds to the upward shift of the black-dashed horizontal line to the S point shown in Fig. 6(b). In this case, the offset Δ(ξ) is close to zero.

4. Conclusions

In conclusion, the ATAS in the IR laser-dressed helium atom is theoretically investigated by calculating single-atom frequency-dependent response function. The main interest is focused on the chirp effect of the attosecond pulse on the generated ATAS. By numerically solving both the fully 3D-TDSE and the three-level model, we demonstrate that although the structure of the absorption spectra is stable against varying chirp rate of the attosecond pulse, the attosecond chirp can induce a temporal shift of the absorption spectrum. We quantitatively identify that the chirp-induced shift distance relative to the chirp-free case is linearly dependent on the attosecond chirp rate and closely related with the IR pulse wavelength. Moreover, it is found that different absorption lines exhibit distinct dependencies on the attosecond chirp. In order to understand these features, we develop an analytical method based on the combination of the time-frequency analysis and the two-photon absorption picture and successfully reproduce the simulation result. Our result reveals that the attosecond chirp effect on ATAS originates from the indirect two-photon absorption processes through a set of the virtual states. This kind of the chirp effect might suggest a route towards the all-optical measurement of the chirp of the attosecond pulse.

Funding

National Natural Science Foundation of China (Grants No. 61690223, No. 11561121002, No. 61521093, No. 11127901, No. 11227902, No. 11404356, No. 11574332); the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDB16); Shanghai Commission of Science and Technology Sailing Project (Grant No. 14YF1406000); Shanghai Institute of Optics and Fine Mechanics Specialized Research Fund (Grant No. 1401561J00); and Youth Innovation Promotion Association CAS.

References and links

1. H. Niikura, F. Légaré, R. Hasbani, A. D. Bandrauk, M. Y. Ivanov, D. M. Villeneuve, and P. B. Corkum, “Sub-laser-cycle electron pulses for probing molecular dynamics,” Nature 417(6892), 917–922 (2002). [CrossRef]   [PubMed]  

2. E. Goulielmakis, V. S. Yakovlev, A. L. Cavalieri, M. Uiberacker, V. Pervak, A. Apolonski, R. Kienberger, U. Kleineberg, and F. Krausz, “Attosecond Control and measurement: lightwave electronics,” Science 317(5839), 769–775 (2007). [CrossRef]   [PubMed]  

3. Z.-H. Loh, M. Khalil, R. E. Correa, R. Santra, C. Buth, and S. R. Leone, “Quantum state-resolved probing of strong-field-ionized xenon atoms using femtosecond high-order harmonic transient absorption spectroscopy,” Phys. Rev. Lett. 98(14), 143601 (2007). [CrossRef]   [PubMed]  

4. M. Kovačev, S. V. Fomichev, E. Priori, Y. Mairesse, H. Merdji, P. Monchicourt, P. Breger, J. Norin, A. Persson, A. L’Huillier, C.-G. Wahlström, B. Carré, and P. Salières, “Extreme Ultraviolet Fourier-Transform Spectroscopy with High Order Harmonics,” Phys. Rev. Lett. 95(22), 223903 (2005). [CrossRef]   [PubMed]  

5. H. Wang, M. Chini, S. Chen, C.-H. Zhang, F. He, Y. Cheng, Y. Wu, U. Thumm, and Z. Chang, “Attosecond Time-resolved autoionization of argon,” Phys. Rev. Lett. 105(14), 143002 (2010). [CrossRef]   [PubMed]  

6. E. Goulielmakis, Z. H. Loh, A. Wirth, R. Santra, N. Rohringer, V. S. Yakovlev, S. Zherebtsov, T. Pfeifer, A. M. Azzeer, M. F. Kling, S. R. Leone, and F. Krausz, “Real-time observation of valence electron motion,” Nature 466(7307), 739–743 (2010). [CrossRef]   [PubMed]  

7. M. Holler, F. Schapper, L. Gallmann, and U. Keller, “Attosecond Electron Wave-Packet Interference Observed by Transient Absorption,” Phys. Rev. Lett. 106(12), 123601 (2011). [CrossRef]   [PubMed]  

8. M. Chini, B. Zhao, H. Wang, Y. Cheng, S. X. Hu, and Z. Chang, “Subcycle ac Stark Shift of Helium Excited States Probed with Isolated Attosecond Pulses,” Phys. Rev. Lett. 109(7), 073601 (2012). [CrossRef]   [PubMed]  

9. Y. Cheng, M. Chini, X. Wang, A. González-Castrillo, A. Palacios, L. Argenti, F. Martín, and Z. Chang, “Reconstruction of an excited-state molecular wave packet with attosecond transient absorption spectroscopy,” Phys. Rev. A 94(2), 023403 (2016). [CrossRef]  

10. M. Schultze, K. Ramasesha, C. D. Pemmaraju, S. A. Sato, D. Whitmore, A. Gandman, J. S. Prell, L. J. Borja, D. Prendergast, K. Yabana, D. M. Neumark, and S. R. Leone, “Attosecond band-gap dynamics in silicon,” Science 346(6215), 1348–1352 (2014). [CrossRef]   [PubMed]  

11. X. Wang, M. Chini, Y. Cheng, Y. Wu, X.-M. Tong, and Z. Chang, “Subcycle laser control and quantum interferences in attosecond photoabsorption of neon,” Phys. Rev. A 87(6), 063413 (2013). [CrossRef]  

12. S. Chen, M. J. Bell, A. R. Beck, H. Mashiko, M. Wu, A. N. Pfeiffer, M. B. Gaarde, D. M. Neumark, S. R. Leone, and K. J. Schafer, “Light-induced states in attosecond transient absorption spectra of laser-dressed helium,” Phys. Rev. A 86(6), 063408 (2012). [CrossRef]  

13. B. Bernhardt, A. R. Beck, X. Li, E. R. Warrick, M. J. Bell, D. J. Haxton, C. W. McCurdy, D. M. Neumark, and S. R. Leone, “High-spectral-resolution attosecond absorption spectroscopy of autoionization in xenon,” Phys. Rev. A 89(2), 023408 (2014). [CrossRef]  

14. C. Ott, A. Kaldun, L. Argenti, P. Raith, K. Meyer, M. Laux, Y. Zhang, A. Blättermann, S. Hagstotz, T. Ding, R. Heck, J. Madroñero, F. Martín, and T. Pfeifer, “Reconstruction and control of a time-dependent two-electron wave packet,” Nature 516(7531), 374–378 (2014). [CrossRef]   [PubMed]  

15. M. Reduzzi, J. Hummert, A. Dubrouil, F. Calegari, M. Nisoli, F. Frassetto, L. Poletto, S. Chen, M. Wu, M. B. Gaarde, K. Schafer, and G. Sansone, “Polarization control of absorption of virtual dressed states in helium,” Phys. Rev. A 92(3), 033408 (2015). [CrossRef]  

16. M. Lucchini, S. A. Sato, A. Ludwig, J. Herrmann, M. Volkov, L. Kasmi, Y. Shinohara, K. Yabana, L. Gallmann, and U. Keller, “Attosecond dynamical Franz-Keldysh effect in polycrystalline diamond,” Science 353(6302), 916–919 (2016). [CrossRef]   [PubMed]  

17. M. B. Gaarde, C. Buth, J. L. Tate, and K. J. Schafer, “Transient absorption and reshaping of ultrafast XUV light by laser-dressed helium,” Phys. Rev. A 83(1), 013419 (2011). [CrossRef]  

18. J. E. Bækhøj, L. Yue, and L. B. Madsen, “Nuclear-motion effects in attosecond transient-absorption spectroscopy of molecules,” Phys. Rev. A 91(4), 043408 (2015). [CrossRef]  

19. M. Chini, X. Wang, Y. Cheng, and Z. Chang, “Resonance effects and quantum beats in attosecond transient absorption of helium,” J. Phys. At. Mol. Opt. Phys. 47(12), 124009 (2014). [CrossRef]  

20. C. Buth and R. Santra, “Theory of x-ray absorption by laser-dressed atoms,” Phys. Rev. A 75(3), 033412 (2007). [CrossRef]  

21. R. Santra, V. S. Yakovlev, T. Pfeifer, and Z.-H. Loh, “Theory of attosecond transient absorption spectroscopy of strong-field-generated ions,” Phys. Rev. A 83(3), 033405 (2011). [CrossRef]  

22. S. Pabst, A. Sytcheva, A. Moulet, A. Wirth, E. Goulielmakis, and R. Santra, “Theory of attosecond transient-absorption spectroscopy of krypton for overlapping pump and probe pulses,” Phys. Rev. A 86(6), 063411 (2012). [CrossRef]  

23. S. Chen, M. Wu, M. B. Gaarde, and K. J. Schafer, “Quantum interference in attosecond transient absorption of laser-dressed helium atoms,” Phys. Rev. A 87(3), 033408 (2013). [CrossRef]  

24. Z. Q. Yang, D. F. Ye, T. Ding, T. Pfeifer, and L. B. Fu, “Attosecond XUV absorption spectroscopy of doubly excited states in helium atoms dressed by a time-delayed femtosecond infrared laser,” Phys. Rev. A 91(1), 013414 (2015). [CrossRef]  

25. L. Argenti, Á. Jiménez-Galán, C. Marante, C. Ott, T. Pfeifer, and F. Martín, “Dressing effects in the attosecond transient absorption spectra of doubly excited states in helium,” Phys. Rev. A 91(6), 061403 (2015). [CrossRef]  

26. T. Sekikawa, T. Ohno, T. Yamazaki, Y. Nabekawa, and S. Watanabe, “Pulse Compression of a High-Order Harmonic by Compensating the Atomic Dipole Phase,” Phys. Rev. Lett. 83(13), 2564–2567 (1999). [CrossRef]  

27. S. Laulan, J. Haché, H. S. Ba, and S. Barmaki, “Ionization Process of Atoms by Intense Femtosecond Chirped Laser Pulses,” J. Mod. Phys. 4(10), 20–26 (2013). [CrossRef]  

28. S. Barmaki, P. Lanteigne, and S. Laulan, “Control of two-photon double ionization of helium with intense chirped attosecond laser pulses,” Phys. Rev. A 89(6), 063406 (2014). [CrossRef]  

29. M. Chini, X. Wang, Y. Cheng, Y. Wu, D. Zhao, D. A. Telnov, S.-I. Chu, and Z. Chang, “Sub-cycle Oscillations in Virtual States Brought to Light,” Sci. Rep. 3, 1105 (2013). [CrossRef]  

30. C. Liu, Z. Zeng, R. Li, Z. Xu, and M. Nisoli, “Mapping the spectral phase of isolated attosecond pulses by extreme-ultraviolet emission spectrum,” Opt. Express 23(8), 9858–9869 (2015). [CrossRef]   [PubMed]  

31. C. Liu, Z. Zeng, R. Li, Z. Xu, and M. Nisoli, “Attosecond photoionization for reconstruction of bound-electron wave packets,” Phys. Rev. A 90(1), 013403 (2014). [CrossRef]  

32. C. Liu and M. Nisoli, “Complete characterization of a coherent superposition of atomic states by asymmetric attosecond photoionization,” Phys. Rev. A 85(5), 053423 (2012). [CrossRef]  

33. M. Wu, S. Chen, S. Camp, K. J. Schafer, and M. B. Gaarde, “Theory of Strong-field attosecond transient absorption,” J. Phys. At. Mol. Opt. Phys. 49(6), 062003 (2016). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1
Fig. 1 Calculated single-atom response function S ˜ (ω, t D ) using the fully 3D-TDSE model upon scanning the relative delay between the IR pulse and the attosecond pulse with three different chirp rates: (a) ξ=4 , (b) ξ=0 and (c) ξ=4 .
Fig. 2
Fig. 2 Delay-dependent absorption signals near (a) 2p (21.05 eV), (b) 4p (23.73 eV) and (c) the low-energy continuum state (25.55 eV), respectively for three different attosecond chirp rates: ξ=4 (red-solid line), ξ=0 (blue-dashed line) and ξ=4 (black-dot line).
Fig. 3
Fig. 3 Absorption signals near the 2p state (21.05 eV) in the case of three different chirp rates: ξ=4 (red-solid line), ξ=0 (blue-dashed line) and ξ=4 (black-dot line), calculated by three-level model. Other simulation parameters are the same as Fig. 1.
Fig. 4
Fig. 4 The IR wavelength-dependent offset Δ(ξ)= t D (ξ) t D (0) in the case of five different chirp rates: ξ=4 (red-pluses), ξ=2 (green-stars), ξ=0 (blue-circles), ξ=2 (cyan-crosses) and ξ=4 (black-squares) presents the fully 3D-TDSE results. Other parameters are consistent with Fig. 3.
Fig. 5
Fig. 5 Schematic of the three-level model including the 1s , 2p and 3s states. S and S + are eigenstates for the 2p and 3s states dressed by the intense IR pulse. The indirect two-photon pathway from the ground state 1s to the final state S through the virtual state S + is also indicated.
Fig. 6
Fig. 6 (a) Time-frequency spectrogram of the XUV pulses at different chirp rates: ξ=4, 2, 0, 2, 4 respectively. (b) Time-dependent photon energy which has the maximal intensity, extracted from panel (a) for five different chirp rates.
Fig. 7
Fig. 7 The IR wavelength-dependent offset Δ(ξ) for five different chirp rates: ξ=4, 2, 0, 2, 4 calculated by an analytical model based on the two-photon absorption process. Other parameters are consistent with Fig. 4.

Equations (9)

Equations on this page are rendered with MathJax. Learn more.

i t Ψ( r ,t)=[ 1 2 1 r 2 r + l ^ 2 2 r 2 +V(r)+ V I ( r ,t) ]Ψ( r ,t),
V(r)= 1.0+1.231 e 0.662r 1.325r e 1.236r 0.231 e 0.48r r ,
A X (t t D )=Re( i 1 ω X I X 1iξ exp{ i[ ω X ( t t D )+ φ X ]2ln2 ( t t D ) 2 τ X 2 ( 1iξ ) } ),
A IR (t)=Re( i 1 ω IR I IR exp{ i[ ω IR t+ φ IR ]2ln2 t 2 τ IR 2 } ),
S ˜ (ω, t D )=2Im[ d ˜ (ω, t D ) ε ˜ * (ω, t D )], ω>0,
E = E 2p + E 3s ( E 2p E 3s ) 2 +4 d 23 2 E IR 2 (t) 2 ,
E + = E 2p + E 3s + ( E 2p E 3s ) 2 +4 d 23 2 E IR 2 (t) 2 ,
η = 2 d 23 2 E IR 2 (t) ( E 2p E 3s ) 2 +4 d 23 2 E IR 2 (t)( E 2p E 3s ) ( E 2p E 3s ) 2 +4 d 23 2 E IR 2 (t) ,
η + = 2 d 23 2 E IR 2 (t) ( E 2p E 3s ) 2 +4 d 23 2 E IR 2 (t)+( E 2p E 3s ) ( E 2p E 3s ) 2 +4 d 23 2 E IR 2 (t) .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.